Stereoselective Formation of Seven-Coordinate Titanium (IV

crystallizes as an oxygen-bridged dimer in the triclinic space group P-1 with a ) 12.521(6) ... Each seven-coordinate, pentagonal-bipyramidal TiIV has...
0 downloads 0 Views 143KB Size
206

Inorg. Chem. 2000, 39, 206-215

Stereoselective Formation of Seven-Coordinate Titanium(IV) Monomer and Dimer Complexes of Ethylenebis(o-hydroxyphenyl)glycine Maolin Guo, Hongzhe Sun, Shailja Bihari, John A. Parkinson, Robert O. Gould, Simon Parsons, and Peter J. Sadler* Department of Chemistry, University of Edinburgh, King’s Buildings, West Mains Road, Edinburgh EH9 3JJ, U.K. ReceiVed June 10, 1999

Reactions between the antitumor agent titanocene dichloride (Cp2TiCl2) and the hexadentate ligand N,N′-ethylenebis(o-hydroxyphenylglycine) (H4ehpg) have been investigated in aqueous solution and the solid state. The racemic ligands give crystals of the monomer [Ti(ehpg)(H2O)]‚(11/3)H2O (1), while the meso ligand gives the oxo-bridged dimer [{Ti(Hehpg)(H2O)}2O]‚13H2O (2). Complex 1 crystallizes in the monoclinic space group C2/c with a ) 24.149(4) Å, b ) 14.143(3) Å, c ) 19.487(3) Å, β ) 105.371(13)°, V ) 6417.7(19) Å3, Z ) 12, and R(F) ) 0.0499 for 4428 independent reflections having I > 2σ(I), and contains seven-coordinate pentagonal-bipyramidal TiIV with two axial phenolate ligands (Ti-O, 1.869(2) Å). The pentagonal plane contains the two N-atoms at 2.210(2) Å, two carboxylate O-atoms at 2.061(2) Å, and a water molecule (Ti-OH2, 2.091(3) Å). Complex 2 crystallizes as an oxygen-bridged dimer in the triclinic space group P-1 with a ) 12.521(6) Å, b ) 14.085(7) Å, c ) 16.635(8) Å, R ) 80.93(2)°, β ) 69.23(2)°, γ ) 64.33(2)°, V ) 2472(2) Å3, Z ) 4, and R(F) ) 0.0580 for 5956 independent reflections having I > 2σ(I). Each seven-coordinate, pentagonal-bipyramidal TiIV has a bridging oxide and a phenolate as axial ligands. The pentagonal plane donors are H2O, two carboxylate O-atoms, and two NH groups, which form H-bonds to O-atoms both in the same half-molecule (O‚‚‚N, 2.93-3.13 Å) and in the other half-molecule (O‚‚‚N, 2.73-2.75 Å); the second phenoxyl group of each Hehpg ligand is protonated and not coordinated to TiIV, but H-bonds to a nearby amine proton (O‚‚‚N, 2.73-2.75 Å) from the same ligand and to a nearby H2O (O‚‚‚O, 2.68 Å). In contrast to all previously reported crystalline metal-EHPG complexes containing racemic ligands, in which the N(S,S)C(R,R) or N(R,R)C(S,S) form is present, complex 1 unexpectedly contains the N(S,S)C(S,S) and N(R,R)C(R,R) forms. This is attributed to the presence of ring strain in sevencoordinate TiIV complexes. Moreover, the rac ligands selectively form crystals of monomeric 1, while the meso ligand selectively forms crystals of the dimer 2 (N(R,R)C(R,S) or N(S,S)C(S,R)). Complexes 1 and 2 exhibit phenolate-to-TiIV charge-transfer bands near 387 nm, and 2D NMR studies indicate that the structures of 1 and 2 in solution are similar to those in the solid state. Complex 1 is stable over the pH range 1.0-7.0, while 2 is stable only between pH 2.5 and pH 5.5. Cp2TiCl2 reacts with EHPG at pH* 7.0 to give complex 1 with a t1/2 of ca. 50 min (298 K), but complex 2 was not formed at this pH* value. At pH* 3.7, the reaction is very slow: 1 forms with a half-life of ca. 2.5 d, and 2 after ca. 1 week at ambient temperature. The relevance of these data to the possible role of serum transferrin as a mediator for the delivery of TiIV to tumor cells is discussed.

Introduction Biscyclopentadienyl-TiIV complexes and bis(β-diketonato)TiIV complexes have been shown to exhibit low toxicity and high antitumor activities against a wide range of murine and human tumors.1-3 Two of them, titanocene dichloride (Cp2TiCl2) and Budotitane [Ti(bzac)2(OEt)2] (Hbzac ) 1-phenylbutane1,3-dionato), are currently on clinical trial. The initial results with Cp2TiCl2 suggest a lack of cross-reactivity with cisplatin and patterns of antitumour activity different from those produced by platinum-based anticancer drugs.4 Recent reports5 have * To whom correspondence should be addressed. Fax: +44 131 650 6452. E-mail: [email protected]. (1) (a) Ko¨pf-Maier, P.; Ko¨pf, H. Chem. ReV. 1987, 87, 1137-1153 and references therein. (b) Ko¨pf-Maier, P.; Ko¨pf, H. Drugs Future 1986, 11, 297-319 and references therein. (c) Ko¨pf-Maier, P.; Ko¨pf, H. In Metal compounds in Cancer Therapy; Fricker, S. P., Ed.; Chapman & Hall: London, 1994; pp 109-146. (2) Keppler, B. K.; Friesen, C.; Vongerichten, H.; Vogel, E. In Metal Complexes in Cancer Chemotherapy; Keppler, B. K., Ed.; VCH: Weinheim, 1993; pp 297-323. (3) Sadler, P. J. AdV. Inorg. Chem. 1991, 36, 1-48.

shown that Cp2TiCl2 significantly overcomes cisplatin resistance in ovarian carcinoma cell lines. The potent antitumor activity and low toxicity of this compound, supported by the encouraging results from two recent phase I clinical trials6 (one in Germany, one in the U.K.), suggest that titanocene dichloride could be a promising novel chemotherapeutic agent.5 In addition, titanocene dichloride exhibits pronounced antiviral, antiinflammatory, and insecticidal activities.7 Additional medical interest in titanium (4) (a) Berdel, W. E.; Schmoll, H. J.; Scheulen, M. E.; Korfel, A.; Knoche, M. F.; Harstrci, A.; Bach, F.; Baumgart, J.; Sass, G. J. Cancer Clin. Oncol. 1994, 120S, R172. (b) Ko¨pf-Maier, P. Eur. J. Clin. Pharmacol. 1994, 47, 1-14. (c) Moebus, V. J., Stein, R.; Kieback, D. G.; Runnebaum, I. B.; Sass, G.; Kreienberg, R. Anticancer Res. 1997, 17, 815-821. (5) Christodoulou, C. V.; Eliopoulos, A. G.; Young, L. S.; Hodgkins, L.; Ferry, D. R.; Kerr, D. J. Br. J. Cancer 1998, 77, 2088-2097. (6) (a) Kortel, A.; Schmol, H. J.; Scheulen M. E.; Gru¨ndel. O.; Harstrick. A.; Knoche. M.; Fels, L. M.; Bach, F.; Baugmart, J.; Sass, G.; Thiel, E.; Berdel, W. E. ASCO 1996, 356, 38. (b) Christodoulou, C. V.; Ferry, D. R.; Fyfe, D. W.; Young, A.; Doran, J.; Sheehan, T. M. T.; Eliopoulos, A. G.; Hale, K.; Baumgart, J.; Sass, G.; Kerr, D. J. J. Clin. Oncol. 1998, 16, 2761-2769.

10.1021/ic990669a CCC: $19.00 © 2000 American Chemical Society Published on Web 12/29/1999

Formation of Seven-Coordinate TiIV Complexes arises from the recent use of 45Ti in radiopharmaceuticals.8 However, in contrast to platinum-based anticancer drugs,9 very little is known about the biological chemistry of titanium compounds. The cytotoxic action of titanium complexes, and their vanadium and molybdenum analogues, may involve mechanisms different from those of cisplatin.2,10-13 Formation of TiIV-DNA complexes is thought to be important for the antitumor properties of Cp2TiCl2, which inhibits DNA synthesis rather than RNA and protein synthesis, and from which titanium accumulates in nucleic-acid-rich regions of tumor cells after in vivo or in vitro administration.14-16 However, unlike cisplatin, Cp2TiCl2 does not bind to DNA bases strongly at physiological pH, but forms strong complexes with nucleotides only at low pH.17 Also, there is no evidence for stable complexes of the V or Mo analogues with nucleotides or DNA under physiological conditions.10a,13c This raises doubts about DNA as the predominant target.10a,13c Efforts to identify the biologically active Ti species have been largely unsuccessful due to the rapid hydrolysis of the titanium(IV) complexes at neutral pH, which results in precipitation of uncharacterized polymeric hydrolysis products.2,11a Transferrin is an 80 kDa iron-transport protein present in the serum of vertebrates at a concentration of about 35 µM.18 It carries iron(III) in blood at pH 7.4, delivers it to cells via receptor-mediated endocytosis, and releases it at pH ca. 5-5.5 in endosomes.18,19 In human serum, only 30% of the transferrin sites are saturated with iron. Many other metal ions have been reported to bind reversibly to human serum transferrin.18 Moreover, transferrin has been implicated in transport and delivery of therapeutic metal ions such as 67GaIII and RuIII to cancer cells.20,21 Since we have recently found that TiIV forms a strong complex with human serum transferrin (hTF) by binding (7) Ko¨pf-Maier, P.; Ko¨pf, H. Struct. Bonding 1988, 70, 105-185. (8) Ishiwata, K.; Ido, T.; Monma, M.; Murakami, H.; Fukuda, M.; Kammeyama, K.; Yamada, K.; Endo, S.; Yoshioka, S.; Sato, T.; Matsuzawa, T. Appl. Radiat. Isot. 1991, 42, 707-712. (9) (a) Sudquist, W. I.; Lippard, S. J. Coord. Chem. ReV. 1990, 100, 293322. (b) Berner-Price, S, J.; Sadler, P. J. Coord. Chem. ReV. 1994, 151, 1-40. (c) Reedijk, J. Chem. Commun. 1996, 801-806. (d) Guo, Z.; Sadler, P. J.; Angew. Chem., Int. Ed. Engl. 1999, 38, 4001-4019 and reference therein. (10) (a) Kuo, L. Y.; Liu, A. H.; Marks, T. J. Met. Ions Biol. Syst. 1996, 33, 53-85, and references therein. (b) Yang, P.; Guo, M. Met.-Based Drugs 1998, 5, 41-58. (c) Yang, P.; Guo, M. Coord. Chem. ReV. 1999, 185-186, 189-211 and references therein. (11) (a) Toney, J. H.; Marks, T. J. J. Am. Chem. Soc. 1985, 107, 947953. (b) Murray, J. H.; Harding, M. M. J. Med. Chem. 1994, 37, 1936. (c) Yang, P.; Guo, M. Met.-Based Drugs 1998, 5, 146-153. (d) Mokdsi, G.; Harding, M. M. Met.-Based Drugs 1998, 5, 207-215. (12) Toney, J. H.; Brock, C. P.; Marks, T. J. J. Am. Chem. Soc. 1986, 108, 7263-7274. (13) (a) Kuo, L. Y.; Mercouri, G.; Kanatzidis, M. S.; Marks, T. J. J. Am. Chem. Soc. 1987, 109, 7207. (b) Kuo, L. Y.; Kanatzidis, M. S.; Sabat, M.; Tipton, L.; Marks, T. J. J. Am. Chem. Soc. 1991, 113, 90279045. (c) Harding, M. M.; Mokdsi, G.; Mackay, J. P.; Prodigalidad, M.; Lucas, S. W. Inorg. Chem. 1998, 37, 2432-2437. (d) McLaughlin, M. L.; Cronan, J. M., Jr.; Schaller, T. R.; Sneller, R. D. J. Am. Chem. Soc. 1990, 112, 8949-8952. (14) Ko¨pf-Maier, P.; Ko¨pf, H.; Wagner, W. Cancer Chemother. Pharmacol. 1981, 5, 237. (15) Ko¨pf-Maier, P.; Krahl, D. Chem.-Biol. Interact. 1983, 44, 317-328. (16) (a) Ko¨pf-Maier, P.; Martin, R. Virch. Arch. B Cell Pathol. 1989, 57, 213-222. (b) Ko¨pf-Maier, P. J. Struct. Biol. 1990, 105, 35-45. (17) Guo, M.; Sadler, P. J. Unpublished results. (18) (a) Baker, E. N. AdV. Inorg. Chem. 1994, 41, 389-463. (b) Aisen, P. Met. Ions Biol. Syst. 1998, 35, 585-631. (c) Sun, H.; Cox, M. C.; Li, H.; Sadler, P. J. Struct. Bonding 1997, 88, 71-102. (d) Sun, H.; Li, H.; Sadler, P. J. Chem. ReV. 1999, 99, 2817-2842. (19) (a) Weaver, J.; Pollack, S. Biochem. J. 1989, 261, 789-792. (b) Weaver, J.; Zhan, H.; Pollack, S. Br. J. Haematol. 1993, 83, 138144. (20) Chitambar, C. R.; Matthaeus, W. G.; Antholine, W. E.; Graff, K.; O’Brien, W. J. Blood 1988, 72, 1930.

Inorganic Chemistry, Vol. 39, No. 2, 2000 207 Chart 1. Schematic Drawing of N,N′-Ethylenebis[(o-hydroxyphenyl)glycine] (H4ehpg)a

a The asterisks indicate the chiral R-carbons. The labeling scheme indicated is used for protons in NMR spectra. For complex 2, a dash is used to label atoms on the uncoordinated phenol (H3′, H4′, etc.)

to the specific FeIII binding sites,22 we have now studied in detail the binding of TiIV to the phenolic ligand ethylenebis(ohydroxyphenylglycine) (EHPG). This ligand itself contains donor groups similar to those of transferrin (2Tyr, His, Asp, and CO32-) and has long been used to model metal-transferrin binding. Metal-EHPG complexes have been shown to model not only many of the physical properties of metal-transferrin complexes but also a number of other iron-phenolate proteins.23-34 In addition, EHPG has been suggested as a potential radiopharmaceutical carrier for positron emission tomography.33 We have studied the interaction of EHPG with titanocene dichloride and titanium(IV) citrate in aqueous solution over a wide range of pH values, and here report the NMR studies and X-ray crystal structures of two novel Ti-EHPG complexes. The structures reveal a unique stereoselective recognition of the EHPG ligands by TiIV. Experimental Section Materials. The ligand H4ehpg (see Chart 1) was purchased from Sigma Chemical Co. and purified as described previously.26 No attempt was made to separate the racemic R,R/S,S and mesomeric R,S isomers of the ligand, which are present in an unspecified mixture in commercial preparations (subsequently determined to be ca. 1:1 in our samples). Cp2TiCl2 was purchased from the Arcos Chemical Co., and monosodium citrate from the Aldrich Chemical Co. Titanium(IV) citrate was prepared as described previously.22 All other chemicals were AR grade and were used as received. NMR Spectroscopy. 1H (500 or 600 MHz) NMR spectra were recorded in D2O on a Bruker DMX 500 or Varian 600 NMR (21) Kratz, F.; Hartmann, M.; Keppler, B. K.; Messori, L. J. Biol. Chem. 1994, 269, 2581-2588. (22) Sun, H.; Li, H.; Weir, R.; Sadler, P. J. Angew. Chem., Int. Ed. Engl. 1998, 37, 1577-1579. (23) Gaber, B. P.; Miskowski, V.; Spiro, T. G. J. Am. Chem. Soc. 1974, 96, 6868. (24) Ainscough, E. W.; Brodie, A. M.; Plowman, J. E.; Brown, K. L.; Addison, A. W.; Gainsford, A. R. Inorg. Chem. 1980, 19, 9, 3655. (25) Harris, W. R.; Carrano, C. J.; Pecoraro, V. L.; Raymond, K. N. J. Am. Chem. Soc. 1981, 103, 2231. (26) Pecoraro, V. L.; Harris, W. R.; Carrano, C. J.; Raymond, K. N. Biochemistry 1981, 20, 7033. (27) Patch, M. G.; Simolo, K. S.; Carrano, C. J. Inorg. Chem. 1982, 21, 2972-2977. (28) Riley, P. E.; Pecoraro, V. L.; Carrano, C. J.; Raymond, K. N. Inorg. Chem. 1983, 22, 3096-3103 (29) Patch, M. G.; Simolo, K. P.; Carrano, C. J. Inorg. Chem. 1983, 22, 2630. (30) Pecoraro, V. L.; Bonadies, J. A.; Marrese, C. A.; Carrano, C. J. J. Am. Chem. Soc. 1984, 106, 3360. (31) Carrano, C. J.; Spartalian, K.; Appa Rao, G. V. N.; Pecoraro, V. L.; Sundaralingam, M. V. J. Am. Chem. Soc. 1985, 107, 1651-1658. (32) Bonadies, J. A.; Carrano, C. J. J. Am. Chem. Soc. 1986, 108, 40884095. (33) Bannochie, C. J.; Martell, A. E. J. Am. Chem. Soc. 1989, 111, 47354742. (34) Lin, W.; Welsh, W. J.; Harris, W. R. Inorg. Chem. 1994, 33, 884890.

208 Inorganic Chemistry, Vol. 39, No. 2, 2000 spectrometer. 1H chemical shifts were referenced to 1,4-dioxane (internal). The 2D COSY, 2D DQFCOSY, and 2D NOESY data were obtained using standard techniques. All spectra were recorded at 298 K unless otherwise stated. The water resonance was suppressed by presaturation or via the WATERGATE pulsed-field-gradient sequence.35 pH Measurements. The pH values were determined using a Corning 240 pH meter equipped with an Aldrich micro combination electrode, calibrated with Aldrich buffer solutions at pH 4, 7, and 10. The pHmeter readings for D2O solutions are recorded as pH* values, i.e., uncorrected for the effects of deuterium on the glass electrode. Adjustments of pH were made with HCl or NaOH (DCl or NaOD for samples in D2O). UV-Vis and FT-IR Spectra. UV-vis spectra at different pH values were recorded in 0.1 M NaCl solution on a Shimadzu UV-2501PC spectrophotometer at 298 K, using 1 cm path length cells. Concentrations of complexes 1 and 2 were determined by measuring the Ti content using the ICP-AES technique. The FT-IR spectra were recorded on a Perkin-Elmer Paragon 1000 spectrometer using KBr pellets. Synthesis of [Ti(ehpg)(H2O)]‚(11/3)H2O (1). Cp2TiCl2 (124.5 mg, 0.5 mmol) was dissolved in boiling distilled water (10 mL), and H4ehpg (180 mg, 0.5 mmol) was added slowly with stirring. The mixture was stirred for 30 min with addition of drops of cold water to keep the volume approximately the same. The resulting deep red-orange solution was filtered while hot to remove traces of unreacted ligand. After cooling to ambient temperature, the pH of the solution was ca. 1.7. One day later, regular rhombic orange crystals were obtained which were suitable for X-ray crystallography. The crystals were filtered and washed 3× with cold water. Yield 73.6% (82.56 mg, based on EHPG(rac), or 36.8% based on Ti). Anal. Calcd for [Ti(C18H16N2O6)(H2O)]‚ 5H2O, C18H18N2O7Ti‚5H2O: C, 42.20; H, 5.50; N, 5.47. Found: C, 42.00; H, 5.32; N, 5.28. FT-IR data (cm-1): 3423br, 1637vs, 1597s, 1478m, 1452m, 1340m, 1250vs, 1188w, 1072w, 1032w, 963w, 901s, 829m, 779m, 761m, 730w, 691w, 646s, 501m, 447w, 428m. Similar crystals of 1 were also obtainable in lower yield by reaction of titanium(IV) citrate with EHPG (1:1, 20 mM each) in water at 277 K, pH 6.1, for several days. Synthesis of [{Ti(Hehpg)(H2O)}2O]‚13H2O (2). The pH of a solution of 200 mL of titanocene dichloride (249.5 mg, 1.0 mmol) containing 1.0 equiv of H4ehpg was adjusted to 3.7, and the solution was left at ambient temperature. A week later, large thin orange rectangular plate crystals were obtained, which were suitable for X-ray crystallography. The crystals were fitered and washed 3× with cold water. Yield 47.4% (117 mg, based on EHPG(meso), or 23.7% based on Ti). Anal. Calcd for C36H38N4O15Ti2‚9H2O: C, 42.21; H, 5.50; N, 5.47. Found: C, 42.53; H, 5.35; N, 5.44. FT-IR data (cm-1): 3422br, 3211br, 1638vs, 1597s, 1480m, 1458m, 1353m, 1294w, 1267m, 1187w, 1106w, 1062w, 1053w, 1006w, 966w, 902w, 790m, 760m, 738m, 702s, 616s, 500m, 434m. Similar crystals of pure 2 were also obtainable in very low yield from solutions of the same concentration as above at pH 1.9 or at pH 5.0 after about a week at ambient temperature. X-ray Crystallography. X-ray crystallographic studies of 1 and 2 were carried out on a Stadi-4 diffractometer equipped with an Oxford Cryosystems low-temperature device. Single crystals were mounted at ambient temperature on the ends of quartz fibers in RS3000 oil, and data were collected at 220 K with Mo KR radiation for 1 and Cu KR for 2, using a θ/ω scan mode. ψ-scan data showed no significant intensity variation for 1, so no absorption correction was applied. The crystal structures were solved by Patterson methods (DIRDIF) (1) and by direct methods (SHELXS) (2). H-atoms attached to C and N were placed in calculated positions, while those attached to O were located in difference Fourier maps; all were allowed to ride on their parent atoms during subsequent refinement. Both structures were refined by full-matrix least-squares against F2 (SHELXL). All non-H-atoms were modeled with anisotropic displacement parameters (adps). The final difference map maximum and minimum were +0.75 and -0.38 e Å-3 (1) and +0.82 and -0.78 e Å-3 (2). Further crystallographic information is summarized in Table 1. (35) Piotto, M.; Saudek, V.; and Sklenar, V. J. Biomol. NMR 1992, 2, 661.

Guo et al. Table 1. Crystal Structure Data for Complexes 1 and 2 chemical formula fw color cryst syst space group a (Å) b (Å) c (Å) R (deg) β (deg) γ (deg) V (Å3) Z T (K) Fcalcd (g‚cm-3) µcalcd (mm-1) F(000) θ range (deg) no. of reflns collected/unique no. of reflns with I > 2σ(I) no. of params R1 (F0 > 2σ(Fo)) wR2 (all data)

1

2

[Ti(C18H16N2O6)(H2O)]‚(11/3)H2O 488.30 deep orange monoclinic C2/c 24.149(4) 14.143(3) 19.487(3) 90 105.371(13) 90 6417.7(19) 12 220(2) 1.516 0.462 3056 2.56-25.03 9463/5662

[Ti2O(C18H17N2O6)2(H2O)2]‚13H2O 1096.47 deep orange triclinic P-1 12.521(6) 14.085(7) 16.635(8) 80.93(2) 69.23(2) 64.33(2) 2472(2) 4 220(2) 1.473 3.582 1152 2.84-70.01 8694/8070

4428

5956

434 0.0499 0.1340

729 0.0580 0.1649

Kinetic Studies of Reactions in Solution. These were carried out at pH* values of 7.0 and 3.7 in D2O. An aliquot of Cp2TiCl2(aq) was introduced into a 5 mm NMR tube charged with EHPG solution (pH* 8.5 or 7.4, respectively) to give a final concentration of 5 mM of each reactant. The pH* values fell to 5.7 or 3.7, respectively, and were readjusted with concentrated NaOD. 1H NMR spectra were recorded immediately, and then subsequently at various time intervals thereafter. There was no measurable change in pH* during the reaction. Similar experiments were performed using titanium(IV) citrate instead of Cp2TiCl2. For kinetic analysis of NMR spectra, relative concentrations of each species were calculated at each time point from peak integrals.

Results Preparations of [Ti(ehpg)(H2O)]‚(11/3)H2O (1) and [{Ti(Hehpg)(H2O)}2O]‚13H2O (2). The reaction of Cp2TiCl2(aq) with commercial EHPG ligand (1 mol equiv) in water afforded orange crystals containing both monomeric [Ti(ehpg)(H2O)] (1) and the oxygen-bridged dimer [{Ti(Hehpg)(H2O)}2O] (2), containing the rac and meso isomers of the ligand, respectively (vide infra). The yields of 1 and 2 depended on the pH, the concentrations of the reactants, and the reaction times. Complex 1 has a higher solubility in water (ca. 5 mM) and was stable at low pH (near pH 1), so a good yield of complex 1 was achieved by carrying out the reaction near boiling temperature for 0.5 h at higher concentrations (ca. 50 mM reactants) at pH ca. 1.7. Then crystals formed after the reaction solution was allowed to stand overnight at ambient temperature. The crude crystals obtained under these conditions sometimes contained trace amounts of the dimer 2, but this was readily removed by washing the crystals with cold water during filtration. Pure crystals of 1 were also obtained from reactions at higher pH (ca. 6.1), using Cp2TiCl2 or titanium(IV) citrate as starting material, but the yields were quite low. Complex 2 is stable only over the pH range of ca. 2.5-5.5. It is less soluble ( 8.0, the LMCT band has nearly disappeared, corresponding to the loss of TiIV-phenolate bonding. Complex 2 was less stable than 1. The LMCT band was observable only between pH 2.5 and pH 6.0. At pH > 7.0, this band disappeared completely. Although the absorbance maximum of the racEHPG complexes is constant at 386 nm throughout the pH range 1.0-6.4, the absorbance maximum of the meso-EHPG complex 2 shifted from 388 nm to a slightly higher wavelength as the

210 Inorganic Chemistry, Vol. 39, No. 2, 2000

Guo et al.

Table 2. Selected Bond Lengths (Å) and Angles (deg) for Complexes 1 and 2 (# ) -x, y, -z + 1/2) complex 1a Ti(11)-O(11) Ti(11)-O(821) Ti(11)-N(91) Ti(11)-O(1W1) O(11)-Ti(11)-O(11#) O(11)-Ti(11)-O(821) O(11)-Ti(11)-O(821#) O(11)-Ti(11)-N(91) O(11)-Ti(11)-N(91#) O(11)-Ti(11)-O(1W1) O(821)-Ti(11)-O(821#) O(821)-Ti(11)-N(91) O(821)-Ti(11)-N(91#) O(821)-Ti(11)-O(1W1) N(91)-Ti(11)-N(91#) N(91)-Ti(11)-O(1W1)

complex 1b 1.869(2) 2.061(2) 2.210(2) 2.091(3)

Distances (Å) Ti(22)-O(12) Ti(22)-O(822) Ti(22)-N(92) Ti(22)-O(2W2)

1.866(2) 2.056(2) 2.197(3) 2.077(2)

171.22(14) 91.64(9) 90.91(9) 82.74(10) 90.14(10) 94.39(7) 146.24(12) 71.16(8) 142.47(9) 73.12(6) 71.91(12) 144.05(6)

Angles (deg) O(12)-Ti(22)-O(202) O(12)-Ti(22)-O(822) O(12)-Ti(22)-O(1332) O(12)-Ti(22)-N(92) O(12)-Ti(22)-N(122) O(12)-Ti(22)-O(2W2) O(822)-Ti(22)-O(1332) O(822)-Ti(22)-N(92) O(822)-Ti(22)-N(122) O(822)-Ti(22)-O(2W2) N(92)-Ti(22)-N(122) N(92)-Ti(22)-O(2W2)

168.24(10) 93.18(10) 90.51(10) 82.22(10) 88.25(10) 95.89(11) 145.92(9) 71.31(9) 143.13(9) 72.59(9) 72.41(9) 143.67(10)

Ti(22)-O(202) Ti(22)-O(1332) Ti(22)-N(122)

1.879(2) 2.059(2) 2.231(3)

O(202)-Ti(22)-O(1332) O(202)-Ti(22)-O(822) O(202)-Ti(22)-N(122) O(202)-Ti(22)-N(92) O(202)-Ti(22)-O(2W2)

91.74(10) 91.41(10) 81.60(10) 89.03(10) 95.81(11)

O(1332)-Ti(22)-N(122) O(1332)-Ti(22)-N(92) O(1332)-Ti(22)-O(2W2)

70.80(9) 142.67(10) 73.33(9)

N(122)-Ti(22)-O(2W2)

143.91(9)

Complex 2 Distances (Å) Ti(1)-O(1) Ti(1)-O(1B) Ti(1)-O(4) Ti(1)-O(3) Ti(1)-O(1T) Ti(1)-N(2) Ti(1)-N(1)

1.853(3) 1.874(3) 2.025(3) 2.093(3) 2.084(3) 2.226(3) 2.241(3)

Ti(2)-O(21) Ti(2)-O(1B) Ti(2)-O(24) Ti(2)-O(23) Ti(2)-O(2T) Ti(2)-N(22) Ti(2)-N(21)

1.944(3) 1.803(3) 2.045(3) 2.097(3) 2.095(3) 2.222(3) 2.233(3)

Angles (deg) O(1)-Ti(1)-O(3) O(1)-Ti(1)-O(4) O(1)-Ti(1)-O(1B) O(1)-Ti(1)-O(1T) O(1)-Ti(1)-N(1) O(1)-Ti(1)-N(2) O(3)-Ti(1)-O(4) O(3)-Ti(1)-O(1B) O(3)-Ti(1)-O(1T) O(3)-Ti(1)-N(1) O(3)-Ti(1)-N(2) O(4)-Ti(1)-O(1B) O(4)-Ti(1)-O(1T) O(4)-Ti(1)-N(1) O(4)-Ti(1)-N(2) O(1B)-Ti(1)-O(1T) O(1B)-Ti(1)-N(1) O(1B)-Ti(1)-N(2) O(1T)-Ti(1)-N(1) O(1T)-Ti(1)-N(2) N(1)-Ti(1)-N(2)

91.59(12) 99.40(12) 164.59(11) 93.69(12) 80.70(12) 85.73(12) 142.98(10) 88.24(12) 71.65(11) 71.32(11) 144.25(11) 90.92(11) 72.19(11) 145.40(11) 72.37(11) 100.86(11) 84.65(12) 85.43(12) 142.33(12) 144.07(12) 73.07(12)

pH was lowered (401 nm at pH 0.7). This shift implies the formation of protonated metal chelate species.32 NMR of Ligands. Commercially available EHPG ligand consists of a mixture of meso and rac isomers. Since we were able to prepare TiIV complexes containing either the pure meso or rac ligand, we have been able to characterize individual isomers of the ligand by NMR. The NMR spectra of free mesoor rac-EHPG ligands were obtained by acidifying the corresponding TiIV complex until a colorless solution was obtained. During this process, tiny amounts of white precipitate, presumably polymeric titanium hydroxide, were formed and subsequently filtered off. Since the 1H NMR spectra are in agreement with formation of the free ligand, as can be seen by comparison with the commercial ligand at low pH (Figure S2 in the Supporting Information), no further attempt was made to remove the metal. 1D 1H NMR data on the EHPG isomers at alkaline pH have been published,27,32 but only the R-protons of the

O(21)-Ti(2)-O(23) O(21)-Ti(2)-O(24) O(21)-Ti(2)-O(1B) O(21)-Ti(2)-O(2T) O(21)-Ti(2)-N(21) O(21)-Ti(2)-N(22) O(23)-Ti(2)-O(24) O(23)-Ti(2)-O(1B) O(23)-Ti(2)-O(2T) O(23)-Ti(2)-N(21) O(23)-Ti(2)-N(22) O(24)-Ti(2)-O(1B) O(24)-Ti(2)-O(2T) O(24)-Ti(2)-N(21) O(24)-Ti(2)-N(22) O(1B)-Ti(2)-O(2T) O(1B)-Ti(2)-N(21) O(1B)-Ti(2)-N(22) O(2T)-Ti(2)-N(21) O(2T)-Ti(2)-N(22) N(21)-Ti(2)-N(22) Ti(1)-O(1B)-Ti(2)

87.53(12) 91.51(12) 167.36(11) 93.77(12) 82.73(12) 86.20(12) 144.09(11) 91.27(12) 72.33(11) 70.53(11) 143.41(12) 96.79(12) 71.92(11) 144.88(11) 72.14(11) 97.85(12) 85.01(12) 87.30(12) 142.81(12) 144.04(12) 72.93(12) 160.22(16)

isomers gave distinguishable signals under these analysis conditions. In addition, resonances for the phenyl ring protons of the isomers were unassigned. At pH* values below 2, resonances for all the protons of the isomers of EHPG are well resolved. The 1H NMR spectrum of the commercial ligand is reproduced exactly by combining the spectra of the individual meso and rac isomers. Unambiguous assignments were established by 2D [1H,1H] COSY and NOESY NMR experiments (Figure S3 in the Supporting Information), and the chemical shifts are listed in Table 3. NMR of Complexes. Solution-state 1H NMR studies of the diamagnetic TiIV complexes 1 and 2 allowed comparison of some of their structural features with those in the solid state. Peak assignments were established by 2D NMR techniques. The 2D COSY and 2D NOESY NMR spectra of 1 and 2 in acidic solution are shown in Figures 4 and 5 and Figure S4 in the Supporting Information. The chemical shifts and assignments

Formation of Seven-Coordinate TiIV Complexes

Inorganic Chemistry, Vol. 39, No. 2, 2000 211

Table 3. Comparison of the1H NMR Chemical Shifts (δ) of Complexes 1 and 2 and the Rac and Meso Isomers of EHPG in D2O at Various pH Values, at 298 Ka pH* R-H methylene phenol H3b [H3′]c H4 [H4′] H5 [H5′] H6 [H6′] a

EHPG (rac)

EHPG (meso)

complex 1 (rac)

complex 2 (meso)

1.38 5.08(s) 3.43(m), 3.21(m) 6.96(d) 7.41(t) 7.00(t) 7.28(d)

1.33 5.11(s) 3.33(m) 6.98(d) 7.43(t) 7.02(t) 7.27(d)

1.37 4.86(s) 3.01(d), 2.66(d) 6.73(d) 7.40(t) 7.08(t) 7.40(d)

3.41 4.94(s), 4.81(s)d 3.55(t), 3.32(d), 2.90(d), 2.65(t) 6.66(d), [6.93(d)] 7.36(t), [7.34(t)] 7.02(t), [6.99(t)] 7.39(d), [7.34(d)]

For atom labels see Chart 1. b Closed phenolate ring. c Uncoordinated phenol. d At 308 K.

Figure 4. 2D NOESY spectrum (800 ms mixing time, 308 K) of the monomer complex 1 in D2O (together with a COSY spectrum, Figure S3), which establishes the NMR resonance assignments. Strong NOE peaks are observed for the phenyl protons, NCH2CH2N protons, R-H to NCH2CH2N, and R-H to H6. For the labeling system, see Chart 1. a, axial protons; e, equatorial protons.

Figure 3. UV-vis spectra of complexes 1 and 2 showing the LMCT bands at various pH values: (A) complex 1, pH 0.73-4.12; (B) complex 1, pH 4.12-8.06; (C) complex 2, pH 0.70-3.37; (D) complex 2, pH 4.30-8.03. Concentration: 1, 3.08 × 10-5 M, 386 ) 9600 M-1 cm-1, pH 4.1; 2, 1.86 × 10-5 M, 388 ) 9480 M-1 cm-1, pH 4.3 (298 K, µ ) 0.1 M, 0.1 M NaCl).

are shown in Table 3, together with data for the meso and rac isomers of the EHPG ligand. Both upfield (low-frequency) and downfield (high-frequency) shifted signals were observed for the complexes compared with the corresponding ligands. The phenolate-TiIV coordination causes a marked upfield shift of the H3 resonance (ca. 0.2 ppm for the rac isomer and 0.3 ppm for the meso isomer), but downfield shifts for H6 (ca. 0.17 ppm for rac and 0.10 ppm for meso), whereas H5 and H4 peaks are only slightly shifted (308 K, the other R-H peak (2H, 4.81 ppm at 308 K) was readily observed. 2D NOESY NMR experiments revealed connectivities between the R-H at 4.94 ppm and the bound phenolate. The ethylenediamine region of the spectrum of the dimer showed two triplets at 3.55 and 2.65 ppm, and two doublets at 3.32 and 2.90 ppm, each of which is further split into fine doublets, representing an AA′BB′ system. This indicates that the HNCH2CH2NH groups are also fixed in a gauche conformation in solution. The assignment of the NMR peaks is shown in Figure 5. The chemical shifts of complexes 1 and 2 were unchanged over the pH* ranges studied (7.5-1.0 and 6.0-1.4, respectively). pH Stability of Complexes 1 and 2. The stability of complexes 1 and 2 at different pH values was further investigated by 1H NMR. Complex 1 is stable over the pH* range 1.0-7.0, but below pH* 1.0, dissociation occurred slowly, the 1H NMR spectrum being consistent with the appearance of the free ligand. However, at pH* g 7.5, many new peaks appeared,

212 Inorganic Chemistry, Vol. 39, No. 2, 2000

Guo et al.

Figure 6. Kinetics of the reactions between Cp2TiCl2(aq) and EHPG in D2O at pH* 7.0 and 298 K followed by 1H NMR, showing the gradual release of the Cp ligands and formation of complex 1. Peak labels: Cp, cyclopentadiene; m and r, meso and rac forms of the free ligand H4ehpg. The same reaction at pH* 3.7 gives rise to both complex 1 and complex 2; see Figure S4.

later. Deep orange crystals were isolated from the solution after 8 d at ambient temperature. Discussion

Figure 5. (A, top) 2D DQFCOSY 1H NMR spectrum (500 MHz, 298 K) and (B, bottom) 2D NOESY spectrum (500 ms mixing time, 308 K) of the dimer complex 2 in D2O, establishing the NMR resonance assignments. Strong NOE peaks are observed for the phenyl protons, for the geminal NCH2CH2N protons (a/b and c/d, see the inset), and for R-H to H6 for the ring-closed but not for the ring-opened side of the ligand. For the labeling system, see Chart 1.

indicating that 1 decomposes at alkaline pH (Figure S5A in the Supporting Information). Complex 2 is less stable than 1, and only over the pH range 2.5-5.5. At pH* > 5.5, many new 1H NMR resonances appeared (Figure S5B in the Supporting Information), indicating the decomposition of complex 2. No attempt was made to identify all the decomposition products. Reactions in Solution. The interaction between Cp2TiCl2 and EHPG was followed by 1H NMR in D2O at two pH* values, pH* 7.0 and 3.7. At pH* 7.0, the Cp ligands were readily released as can be seen from the decrease in intensity of the 1H NMR peak for TiIV-bound Cp at 6.42 ppm, and the appearance of peaks at 6.64 and 6.57 ppm assignable to free cyclopentadiene (Figure 6).11a Both of the isomers of EHPG bind to TiIV, but the rac ligand binds faster than the meso isomer as can be seen from the rate of decrease of the R-H NMR peaks. The increase in intensity of the two deceptive quasi-quartets at 2.99 and 2.64 ppm (NCH2CH2N), the doublet at 6.72 ppm (H3), the triplet at 7.06 ppm (H5), and the multiplet at 7.39 ppm (H4 and H6) is indicative of the gradual formation of the rac complex 1 with a t1/2 of ca. 50 min. However, no NMR peaks characteristic of the meso complex 2 were observed during the reaction, suggesting that species containing the meso ligand were other than complex 2 at this pH value. At pH* 3.7, Cp2TiCl2 reacted very slowly with both of the EHPG isomers, as can be seen from the NMR spectra (Figure S6 in the Supporting Information). The reaction leading to the formation of the rac complex 1 required about one week for completion (t1/2 of ca. 2.5 d), and weak NMR peaks characteristic of the meso dimer, complex 2, did not appear until 7 d

The yield of monomeric complex [Ti(ehpg)(H2O)]‚(11/3)H2O (1) or dimeric complex [{Ti(Hehpg)(H2O)}2O]‚13H2O (2) depends on pH, concentrations, and the reaction time. Complex 1 has higher solubility in water (ca. 5 mM) than complex 2 (143.5°) to encourage the coordination of a water molecule in the large hole thus formed. Therefore, it is understandable that TiIV becomes sevencoordinate in these circumstances. The VIV-EHPG complex also has long V-N bonds, but V(IV) has a smaller ionic radius (0.63 Å), one of the oxygen ligands (VdO) in the trans O-V-O unit is not connected by a chelate ring (one of the phenolates is not coordinated to VIV in this complex), and the O-V-O angle is only 103°, not large enough to accommodate a water molecule. A similar analysis may also apply to the spherically symmetric high-spin d5 FeIII-EHPG complexes, which also have relatively long Fe-N bond lengths (ca. 2.16 Å) and a medium ionic radius for FeIII (0.64 Å). Although they are six-coordinate in the crystal lattice with O-Fe-O angles of 105-112°, optical spectroscopic properties suggest that FeIII-EHPG may be seven-coordinate in aqueous solution with an additional bound water ligand.29 The most distinguishing feature of these TiIV-EHPG structures is their stereochemistry. The EHPG ligand itself contains two chiral carbon atoms, so it has two diastereomeric forms, a pair of enantiomers (R,R and S,S) and a meso form (R,S or S,R). However, the EHPG ligand contains two prochiral nitrogen atoms, and so the two amine nitrogens become chiral when bound to a metal. Thus, there are four chiral atoms in metalEHPG complexes, resulting in a total of 16 possible configurations. As discussed earlier,34 only a gauche conformation for the NCH2CH2N group in the metal complexes is possible. This constraint and the 2-fold axis in the meso conformers reduce the original 16 possible conformers to only 6. In the three conformers with nitrogen atoms in the N(S,S) configuration, the chiral carbon atoms can be labeled as C(R,R), C(S,S), and C(R,S). If the nitrogen atoms are in the N(R,R) form, a similar argument holds, since they are mirror images of the three N(S,S) conformers. The complex with absolute chirality of N(S,S)C(R,R), which has been designated as (R,R) rac34 or ΛRR28 in some previous reports, and its mirror image isomer N(R,R)C(S,S), ∆SS,28 have the six-membered phenolate chelate rings in the equatorial positions and five-membered carboxylate chelate rings in axial sites. One of the mirror isomers with the

214 Inorganic Chemistry, Vol. 39, No. 2, 2000 Chart 2. Schematic Drawing of Octahedral M-EHPG Isomersa

a

The asterisks indicate chiral atoms.

N(S,S)C(R,R) configuration is shown in Chart 2A. The complex with N(S,S)C(S,S), which has been designated as (S,S) rac34 or ∆RR,28 in the literature, and its mirror isomer N(R,R)C(R,R) both have the six-membered chelate phenolate rings in axial sites. One of the mirror isomers with the N(S,S)C(S,S) configuration is shown in Chart 2B. The N(S,S)C(R,S) meso isomer, which has been designated as (R,S) meso in the literature, and its mirror isomer N(R,R)C(S,R) have one axial and one equatorial phenolate, as shown in Chart 2C. Although X-ray crystal structures containing the N(S,S)C(R,R) and N(R,R)C(S,S) ((R,R) rac, ΛRR and ∆SS) conformers of FeIII, CoIII, GaIII, CuII, and VIV and the (R,S) meso conformer of FeIII are available, the crystal structure of a complex containing the corresponding N(S,S)C(S,S) or N(R,R)C(R,R) ((S,S) rac, ΛSS or ∆RR) conformer has not yet been reported, as far as we are aware. The absence of the N(S,S)C(S,S) or N(R,R)C(R,R) ((S,S) rac, ΛSS or ∆RR) conformers in crystals has previously been ascribed to unfavorable crystal packing,43 or to the inherent instability of the (S,S) rac conformer.28 Molecular mechanics studies34 of FeIII-EHPG complexes show that the most stable conformers are the N(S,S)C(R,R) and N(R,R)C(S,S) ((R,R) rac, ΛRR and ∆SS), both in vacuo and in aqueous solution, while the least stable is the N(S,S)C(S,S) or N(R,R)C(R,R) ((S,S) rac, ΛSS or ∆RR) conformer. Stability constants for the (R,R) rac and (R,S) meso complexes of EHPG with a few metal ions (FeIII, NiII, ZnII, CuII, GaIII, and InIII) have been determined,33 and the results show that all the (R,R) rac conformers are more stable by 2-3 kcal/mol than the (R,S) meso conformer for all the metal ions studied. Thus, the presence of the (S,S) rac conformers (N(S,S)C(S,S) and its mirror isomer N(R,R)C(R,R) as pairs in the unit cell) in crystals of 1 was unexpected. In crystals of the dimer 2, both TiIV-EHPG units are in the (R,S) meso form (N(R,R)C(R,S) for both EHPG ligands with its mirror isomer N(S,S)C(S,R) also present in the unit cell). (43) (a) Bailey, N. A.; Cummins, D.; Mckenzie, E. D.; Worthington, J. M. Inorg. Chim. Acta 1976, 18, L13. (b) Bailey, N. A.; Cummins, D.; Mckenzie, E. D.; Worthington, J. M. Inorg. Chim. Acta 1981, 50, 511.

Guo et al. None of the preparations gave rise to TiIV-EHPG N(S,S)C(R,R) or N(R,R)C(S,S) ((R,R) rac, ΛRR and ∆SS) isomers, which have always been observed for other metals studied so far. For octahedral metal-EHPG complexes, greater stability has been proposed33,34 for (R,R) rac isomers due to the geometric selectivity in the placement of both six-membered chelate rings in the equatorial plane defined by the ethylenediamine ring. In the (S,S) rac conformer, both of the six-membered chelate rings are in the axial position, which is a very strained arrangement. In the case of the (R,S) meso isomer, the only possible arrangement is the one with all coordinating groups cis to one another, that is, one axial phenyl and one equatorial phenyl. This structure is intermediate between the highly strained (S,S) rac arrangement and the most favorable (R,R) rac arrangement. The unusual N(S,S)C(S,S) or N(R,R)C(R,R) ((S,S) rac, ΛSS or ∆RR) chirality of TiIV-EHPG in 1 can be understood by considering the factors leading to metal complex stability. The two major factors are ligand basicity and coordination geometry.33 Experimental results have suggested that the two isomers of EHPG have the same basicities (rac 38.0 and meso 37.9).33 Therefore, the coordination geometry is likely to play a major role. Since the bite angle of a six-membered chelate ring28,44 is about 90-95° and substantially greater than that of a fivemembered chelate ring (ca. 70-80°),28,44 more satisfactory octahedral coordination is achieved with both six-membered rings bound to the metal in the equatorial plane that contains the five-membered ethylenediamine group. This is the case for the (R,R) rac conformation in most of the reported metal-EHPG complexes. For the Ti-EHPG complexes, TiIV is sevencoordinate with pentagonal-bipyrimidal geometry, which will be very strained if any of the six-membered chelate rings bound to the metal are in the pentagonal plane. More satisfactory pentagonal-bipyrimidal coordination is achieved with the sixmembered chelate rings in the axial position, which results in the ring-closed monomer [Ti(ehpg)(H2O)]‚(11/3)H2O (1) crystallizing as the N(S,S)C(S,S) and N(R,R)C(R,R) ((S,S) rac, ΛSS or ∆RR) isomers, and the (R,S) meso ligands forming the ringopened oxo-bridged dimer [{Ti(Hehpg)(H2O)}2O]‚13H2O (2). Titanium(IV) readily forms complexes with EHPG(rac), a chelating amino acid derivative, at neutral pH. This has implications for the mechanism of action of titanium anticancer drugs. TiIV is very “hard” and readily hydrolyzes to form insoluble polymeric species at neutral pH. Both titanocene dichloride and Budotitane undergo rapid and complete hydrolysis to form insoluble polymers at physiological pH.2,11a However, biological experiments reveal that titanium is accumulated in the cellular nucleic-acid-rich regions (mainly the nucleus) after in vivo application of titanium drugs.7 Therefore, biomolecules must be involved in the stabilization of TiIV and its transport. The present work shows that TiIV can bind to EHPG(rac) at neutral pH, and is thought to bind in a manner similar to that of to transferrin.22,45 Both titanium (IV) citrate and Cp2TiCl2(aq) can transfer TiIV to the specific FeIII sites of human apotransferrin at pH 7.4. Transferrin may therefore serve to deliver TiIV from the antitumor drug Cp2TiCl2 or Budotitane to cancer cells, which are known to have a higher density of transferrin receptors than normal cells,46 and may release the metal in endosomes inside cells. This may allow these drugs to target biomolecules such as nucleic acids at low pH, or subsequently to bind to other phenolate proteins. Either of these (44) Bannochie, C. J.; Martel, A. E. Inorg. Chem. 1991, 30, 1385-1392. (45) Guo, M.; Sun, H.; Sadler, P. J. Manuscript in preparation. (46) Wagner, E.; Curiel, D.; Cotten, M. AdV. Drug. DeliVery ReV. 1994, 14, 113.

Formation of Seven-Coordinate TiIV Complexes routes could result in TiIV inhibition of DNA synthesis and mitotic activity. Recently, Marks and co-workers also suggested that enzymes could also be targets for metallocene anticancer drugs.10a Further biological investigations of titanium transferrin or other phenolate proteins are therefore warranted. Acknowledgment. We thank the Committee of ViceChancellors and Principals for an ORS award, the University of Edinburgh for a scholarship (M.G.), and the EPSRC (studentship for S.B.) and BBSRC for support. We are grateful to Dr X. Tan (Edinburgh) for his kind advice on crystallization. Supporting Information Available: Figure S1 [molecular structure and atom-labeling scheme for one of the mirror isomers of complex

Inorganic Chemistry, Vol. 39, No. 2, 2000 215 1b], Figure S2 [500 MHz 1H NMR spectra of commercial EHPG and its rac and meso isomers at acidic pH* values in D2O], Figure S3 [500 MHz [1H,1H] (A) COSY and (B) NOESY (mixing time 1.2 s) spectra of the EHPG(rac) isomer in D2O at 310 K, which establish the NMR resonance assignments], Figure S4 [2D [1H,1H] COSY 1H NMR spectrum (500 MHz, 298 K) of the monomer complex 1 in D2O], Figure S5 [1H NMR spectra of (A) complex 1 and (B) complex 2 at different pH* values], Figure S6 [kinetics of the reactions between Cp2TiCl2(aq) and EHPG in D2O at pH* 3.7 followed by 1H NMR], and Figure S7 [a color picture showing the intramolecular hydrogen bonds in complexes 1 and 2]. This material is available free of charge via the Internet at http:/pubs.acs.org. IC990669A